Mechanistic Insights on Plasmon-Driven Photocatalytic Oxidative

5 days ago - The plasmonic electron oscillations in optically excited metallic nanoparticles result in surface confinement of photon energy over much ...
0 downloads 10 Views 4MB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

Article

Mechanistic Insights on Plasmon-Driven Photocatalytic Oxidative Coupling of Thiophenol Derivatives: Evidence for Steady State Photo-Activated Oxygen Qingfeng Zhang, and Hui Wang J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b00660 • Publication Date (Web): 20 Feb 2018 Downloaded from http://pubs.acs.org on February 21, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Mechanistic Insights on Plasmon-Driven Photocatalytic Oxidative Coupling of Thiophenol Derivatives: Evidence for Steady State Photo-Activated Oxygen

Qingfeng Zhang and Hui Wang*

Department of Chemistry and Biochemistry, University of South Carolina, Columbia, South Carolina 29208, United States

* To whom correspondence should be addressed. Email: [email protected]; Phone: 1-803-777-2203; Fax: 1-803-777-9521.

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 34

ABSTRACT The plasmonic electron oscillations in optically excited metallic nanoparticles result in surface confinement of photon energy over much longer time scales in comparison to the unconfined photons traveling at the speed of light, thereby producing an enormous buildup of photon intensity and highly concentrated energetic hot electrons at the nanoparticle surfaces. While the plasmonic hot electrons can be harnessed to drive unconventional photocatalytic molecular transformations at the nanoparticlemolecule interfaces, a set of fundamentally important issues concerning detailed reaction mechanisms still remain poorly understood. Here we use surface-enhanced Raman scattering (SERS) as a unique time-resolving and molecular finger-printing tool to spectroscopically resolve the complex kinetics and underlying pathways of plasmon-driven oxidative coupling of thiophenol derivatives chemisorbed on the surfaces of optically excited plasmonic Ag nanostructures. A hybrid supra-nanostructure composed of a SiO2 bead densely decorated with Ag nanocubes was used as both the SERS substrate and the plasmonic photocatalyst under near infrared excitations. Through deliberately designed time-resolved single-particle SERS measurements, we have been able to pinpoint the effects of excitation power, local-field enhancement, interfacial oxygen abundance, molecular structures, and photothermal heating on the kinetics and yields of the hot electron-driven oxidative coupling reactions. Our time-resolved SERS results provide compelling experimental evidence for the steady state photo-activated oxygen, revealing that the chemical transformations of thiophenol derivatives rather than the photo-activation of interfacial oxygen constitute the kinetic bottlenecks along the multi-step reaction pathways while the overall reaction rates are dynamically maneuvered by the photo-activated oxygen at its steady state concentrations.

ACS Paragon Plus Environment

2

Page 3 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

INTRODUCTION Optically excited plasmonic nanoparticles can catalyze a unique set of interfacial molecular transformations boosted by the enormous buildup of photon intensity1-4 and high abundance of energetic ballistic electrons, also known as hot electrons,5-10 at the nanoparticle surfaces. Due to lack of coherence, the plasmonic hot carriers are fundamentally different from the excitons in semiconductors, making it possible to drive photochemical reactions along unconventional pathways distinct from those involved in conventional thermal catalytic reactions and semiconductor-based photocatalysis.10-16 It has been demonstrated that plasmonic hot electrons can be efficiently harnessed to drive or enhance intriguing

photocatalytic

reactions,

such

as

light-induced

nanocrystal

growth,16-20

photopolymerization,21,22 dissociation of H2,23,24 water splitting,12,25 nitrothiophenol reduction,26-29 selective oxidation,11,14,30-34 and the Suzuki coupling reactions.35-37 While the generation and migration of hot electrons upon plasmonic excitation and decay are fast photophysical processes over time scales from femtoseconds to picoseconds,8,10,15,31 the plasmon-driven photocatalytic reactions typically occur on drastically longer time scales ranging from a few seconds up to several hours,11,13,22-24,30,31,33 strongly indicating that slow interfacial chemical or photochemical processes, most likely the photoactivation of crucial interfacial species or induced transformations of molecular adsorbates, constitute the kinetic bottlenecks along the multi-step reaction pathways. It still remains an immensely challenging task, however, to fully elucidate the detailed reaction mechanisms largely due to the intrinsic complexity of multiple intriguing photophysical, photochemical, and interfacial processes entangled in the plasmon-driven photocatalytic molecular transformations. A key step in plasmon-driven photocatalysis has been identified to be the photo-activation of crucial species upon injection of plasmonic hot electrons into the unpopulated orbitals of the molecular adsorbates, which selectively activates certain types of chemical bonds and creates highly reactive transient species that dynamically modulate the interfacial molecular transformations.10,13,31 This is best manifested by the photo-activation of triplet oxygen, 3O2, adsorbed on plasmonic nanoparticle ACS Paragon Plus Environment

3

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 34

surfaces, which generates transient doublet 2O2-, a highly active oxidative species that selectively drives a series of aerobic oxidation reactions, such as epoxidation of ethylene,11,14,30,34 oxidation of CO,30 and oxidative coupling of aminothiophenol.38 Although multiple experimental evidences have coherently verified the existence of such photo-activated oxygen,13,30,34,38 what the rate-limiting steps are and how the transient photo-activated oxygen dynamically modulates the reaction kinetics still remain fundamentally intriguing open questions, inspiring us to perform detailed kinetic studies to fully understand the dynamic nature of the photo-activated oxygen and pinpoint the underlying ratedetermining factors using plasmon-driven photocatalytic oxidative coupling of thiophenol derivatives as prototypical model reactions. The formation of 4,4'-dimercaptoazobenzene (DMAB) through photocatalytic oxidative coupling of 4-aminothiophenol (4-ATP) adsorbed on optically exited plasmonic nanoparticles represents an enthralling plasmon-driven paradigm for unconventional synthesis of aromatic azo compounds.28,32,33,38 This photocatalytic reaction was first discovered in 2010 when several Raman peaks signifying the characteristic ag modes of DMAB emerged in the surface-enhanced Raman scattering (SERS) spectra upon laser illumination of 4-ATP chemisorbed on metallic nanostructures.32,33 These characteristic SERS peaks of DMAB, first observed by Osawa and coworkers back in 1994,39 had long been controversially interpreted as the b2 modes of 4-ATP enhanced by interfacial charge transfers until Tian and coworkers convincingly proved that 4-ATP underwent an oxidative coupling process with the aid of photo-activated oxygen to form DMAB during SERS measurements.33,38 Although this plasmondriven coupling reaction has been intensively investigated using a large variety of plasmonic nanostructures as the photocatalysts under a diverse range of reaction conditions,28,32,33,38,40-42 several key aspects regarding the detailed mechanisms still remain unclear. (1) The rate-limiting step needs to be identified, based on which the derivation of an empirical rate law becomes possible. (2) It still remains an open question whether the overall reaction kinetics is dynamically tied to the interfacial abundance of the photo-activated oxygen. (3) The search for optimal plasmonic photocatalysts has ACS Paragon Plus Environment

4

Page 5 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

been conducted in a largely empirical fashion due to lack of quantitative understanding of the detailed correlations between plasmonic field intensity and reaction kinetics. (4) The plasmonic photothermal heating may thermally activate the molecular adsorbates on the nanoparticle surfaces and thereby alter the activation energy barriers of interfacial reactions. How the photothermal heating influences the reaction kinetics and yields is well worthy of further scrutiny. (5) Substitution of the hydrogen atoms in the amine group of 4-ATP with other functional groups may significantly modify the reaction energy landscapes. However, such molecular structural effects still remain unexplored. To shed light on these fundamentally intriguing but intrinsically complex issues, we use SERS as a time-resolving and molecular finger-printing tool to fully resolve the complex kinetics of oxidative coupling of several thiophenol derivatives chemisorbed on plasmonic photocatalysts. Time-resolved SERS provides a unique means to quantitatively correlate the reaction kinetics with the plasmonic characteristics of the photocatalysts, enabling us to pinpoint the effects of several key factors dictating the reaction kinetics and yields, such as excitation laser power, local-field enhancement, the abundance of interfacial oxygen, the molecular structures of thiophenol derivatives, and plasmonic photothermal annealing. As revealed by this work, the rate-limiting step is essentially associated with the interfacial chemical transformations of the thiophenol derivatives while the overall reaction kinetics is dynamically maneuvered by the photo-activated oxygen at its steady state concentrations.

METHODS Core-satellites supra-nanoparticles (SNPs), each of which was composed of a spherical SiO2 core densely decorated with Ag nanocubes (NCs) were used as both the SERS substrates and the plasmonic photocatalysts. The SiO2@Ag NCs SNPs were assembled by electrostatically attaching the Ag NCs onto the surfaces of polydiallyl-dimethylammonium (PDDA)-functionalized SiO2 beads through a layer-by-layer assembly process

43

(see details in Supporting Information). The structures,

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 34

compositions, and surface properties of the as-assembled SNPs were characterized by transmission electron microscopy (TEM), scanning electron microscopy (SEM), energy dispersive spectroscopy (EDS), optical extinction spectroscopy, and ζ-potential measurements (see details in Supporting Information). Sub-monolayers of isolated SiO2@Ag NCs SNPs were immobilized onto poly(4vinylpyridine)-functionalized silicon substrates following a previously reported protocol 44 (see details in Supporting Information). To form self-assembled monolayers of the thiophenol derivatives on the nanoparticle surfaces, the SiO2@Ag NCs SNPs immobilized on silicon substrates were incubated with 500 µM 4-ATP, 4-dimethylamino-thiophenol (4-DMATP), or 4-acetamido-thiophenol (4-AATP) ethanolic solution at room temperature for 24 hours, then thoroughly washed with ethanol and water, and finally dried with N2 gas. Time-resolved SERS spectra were collected using a BaySpec NomadicTM Raman microscope built on an Olympus BX51 reflected optical system under 785 nm laser excitation in the confocal mode (focal area of 2 μm diameter). A 50× dark field objective (NA=0.5, WD=10.6 mm, Olympus LMPLFLN-BD) was used for both Raman signal collection and dark field scattering imaging. The laser beam was focused on individual SiO2@Ag NCs SNPs for Raman spectrum collection. The spectral acquisition time was typically 2 seconds unless mentioned otherwise and the power of excitation laser focused on the samples was adjusted using a neutral density filter and measured using a laser power meter (Newport, Model: PMKIT-05-01). The photocatalytic reactions were carried out in various atmospheres, such as ambient air, pure oxygen, pure nitrogen, and oxygen/nitrogen mixtures, at 1 atmospheric pressure and room temperature. Under each experimental condition, the kinetic measurements were repeated on 10 different SiO2@Ag NCs SNPs, one particle at a time. Normal Raman spectra were collected on solid thin films of 4-ATP, 4-DMATP, and 4-AATP on silicon substrates, respectively.

ACS Paragon Plus Environment

6

Page 7 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

RESULTS AND DISCUSSION Our SERS-based kinetic studies involved the use of deliberately designed, bifunctional SiO2@Ag NCs SNPs as both the SERS substrate and the plasmonic photocatalyst. Monodisperse Ag NCs with edge lengths of 35.9 ± 1.62 nm were synthesized following a protocol developed by Xia and coworkers45 and carboxylate-functionalized spherical SiO2 beads with uniform diameters of 1.00 ± 0.0240 µm were purchased from Bangs Laboratories, Inc. (Figure S1 in Supporting Information). When dispersed in water at neutral pH, colloidal SiO2 beads and Ag NCs were both negatively charged on their surfaces. We used a thin layer of positively charged polymer, PDDA, as the linker between the SiO2 bead and the Ag NCs to construct SiO2@Ag NCs SNPs through an electrostatic layer-by-layer assembly approach43 (Figure 1A). The evolution of the surface charges of the particles during the stepwise assembly process was monitored by ζ-potential measurements (Figure S2 in Supporting Glass Information). The surface of each SiO2 bead was densely decorated with Ag NCs, as clearly shown by the SEM images (Figures 1B-1D) and EDS-based elemental analysis (Figures 1C and Figure S3 in Supporting Information). While colloidal Ag NCs exhibited a narrow plasmon resonance band centered at 415 nm, the SiO2@Ag NCs SNPs displayed a broad spectral feature in the optical extinction spectrum with plasmon resonance red-shifted into the near infrared (Figure 1E) due to strong plasmon coupling between adjacent Ag NCs,46 allowing us to excite the plasmons over a broad spectral range across the visible and near infrared regions to achieve large SERS enhancements as well as high photocatalytic efficiency. In this work, we used a continuous wave 785 nm laser as the excitation source for both SERS and photocatalysis. The SiO2 beads did not exhibit any plasmonic spectral features in the visible and near infrared regions, thereby serving solely as a dielectric substrate supporting the Ag NCs. The plasmonic “hot-spots” located inside the sub-10 nm gaps between neighboring Ag NCs provided enormous field enhancements exploitable not only for enhancing Raman signals but for producing high surface abundance of hot electrons as well. The as-assembled SiO2@Ag NCs SNPs could be immobilized onto poly(4-vinylpyridine)-functionalized silicon ACS Paragon Plus Environment

7

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 34

substrates44,47 to form a submonolayer of well-separated individual particles (Figure S4 in Supporting Information). Thiophenol derivatives, such as 4-ATP, 4-DMATP, and 4-AATP, strongly adsorbed onto the Ag NC surfaces through covalent Ag-S interactions to form self-assembled monolayers while wellpreserving the structural integrity of the SNPs without causing observable detachment of Ag NCs from the silica cores (Figure S5 in Supporting Information). We used a confocal Raman microscope/spectrometer to focus the excitation laser onto 2 μm × 2 μm regions containing individual SiO2@Ag NCs SNPs, which allowed us to resolve the photocatalytic reaction kinetics on one SNP at a time through time-resolved SERS measurements.

Figure 1. Assembly of SiO2@Ag NCs SNPs. (A) Schematic illustration of the assembly of SiO2@Ag NCs SNPs. (B) SEM image of SiO2@Ag NCs SNPs. (C) SEM image and EDS elemental distribution maps (based on the intensities of Ag-L, Si-K, and O-K lines) of a SiO2@Ag NCs SNP. (D) TEM image of a SiO2@Ag NCs SNP. The inset shows the TEM image of the entire particle and the region enclosed by the dash box is shown with a higher magnification. (E) Optical extinction spectra of colloidal SiO2 beads, Ag NCs, and SiO2@Ag NCs SNPs.

ACS Paragon Plus Environment

8

Page 9 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Plasmon-driven photocatalytic reactions are initiated by the optical excitation of collective oscillations of valence electrons, also known as plasmons, in the photocatalysts. Following the optical excitation, the plasmonic electron oscillations can dephase through either radiative elastic photon scattering or nonradiative relaxation pathways, namely Landau damping (generation of hot electrons and holes in the metals) or chemical interface damping (direct electron injection into the molecular adsorbates on the nanoparticle surfaces).31 The radiative decay of plasmons gives rise to tremendously enhanced local electromagnetic fields exploitable for SERS, while the nonradiative plasmon decay involves charge transfers between strongly coupled nanoparticles and molecular adsorbates, triggering photochemical transformations on the surfaces of plasmonic nanostructures.13 As illustrated in Figure 2A, the plasmonic excitation of a Ag nanoparticle creates a nonthermal distribution of energetic hot electrons transiently occupying the empty states above the Fermi level while leaving the hot holes below the Fermi level of Ag (-4.3 eV vs. vacuum). The transfer of hot electrons from Ag to molecular adsorbates becomes possible when energetically favorable band alignments exist between the excited hot electrons and the unoccupied adsorbate states. Upon plasmonic excitation at 785 nm (photon energy of 1.58 eV), the hot electrons were distributed in an energy range insufficient for them to be injected into the lowest unoccupied molecular orbital (LUMO) of 4-ATP chemisorbed on Ag (3.8 eV above the Ag Fermi level).48,49 However, previous density functional theory (DFT) calculations showed that the energy of the antibonding 2π* orbitals of molecular 3O2 adsorbed on Ag {100} facet (the dominant facets on the surfaces of Ag NCs45) matched the Ag Fermi level very well,30 thereby facilitating the transfer of a hot electron from Ag to the 2π* orbitals of 3O2 adsorbate to form 2O2-, a short-lived, photo-activated species that may either rapidly react with other molecular adsorbates, as exemplified by the oxidative coupling of thiophenol derivatives studied in this work, or relax to its ground electronic state after the decay of an electron in the antibonding 2π* orbitals back to Ag. Immediately after the photo-activated 2O2- reacted with chemisorbed 4-ATP to initiate the coupling reaction, an electron might rapidly recombine with a hole in the plasmonically excited Ag NC, ACS Paragon Plus Environment

9

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 34

preventing the Ag NC from being oxidized. After the photocatalytic oxidative coupling reactions, the morphology of Ag NCs and the distribution of Ag NCs on the SiO2 core were both well-preserved (Figure S6 in Supporting Information), indicating that the laser illumination and the photocatalytic reactions did not cause any oxidation or dissolution of the Ag NCs observable within the resolution of SEM. The possibility of forming transient, locally oxide Ag surfaces during the photocatalytic reactions, however, had not been ruled out simply based on the ex situ SEM imaging results. While neither the ground state nor the photo-excited interfacial oxygen was spectroscopically resolvable by SERS, the formation of the product, DMAB, could be monitored in real time based on the temporal evolution of the characteristic SERS peaks of DMAB. Although this photocatalytic reaction involved multiple steps many of which were indecipherable by SERS, a simplified version depicted as the dimerization of 4-ATP to form DMAB on Ag surface (Figure 2B) fully captured the essential SERS-resolvable chemical transformation. Figures 2C and 2D show the temporal evolution of SERS spectra of 4-ATP adsorbed on a SiO2@Ag NCs SNP upon exposure to the continuous wave excitation laser (wavelength: 785 nm; power: 0.45 mW; focal plane: 3.14 × 10-8 cm2; power density: 14 kW cm-2) under the ambient aerobic condition at room temperature. Prior to the formation of DMAB, the two major SERS peaks at 1078 and 1595 cm-1 were assigned to the C-S stretching mode and benzene ring mode of 4-ATP, respectively.32,33 As the photocatalytic reaction proceeded, several newly emerged peaks at 1140, 1390, and 1440 cm-1 gradually became more intense, signifying the formation of DMAB.33,38 Meanwhile, the benzene ring mode down-shifted from 1595 to 1575 cm-1, while the C-S stretching mode slightly downshifted by 4 cm-1 upon the formation of DMAB. The transformation of 4-ATP into DMAB also resulted in significantly stronger SERS signals because DMAB had much larger Raman cross-sections than those of 4-ATP.33,38 The detailed peak assignments were listed in Tables S1 and S2 in Supporting Information.

ACS Paragon Plus Environment

10

Page 11 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 2. Kinetics of plasmon-driven oxidative coupling of 4-ATP. (A) Schematic illustration of plasmonic hot electron-driven photo-activation of O2 adsorbed on Ag surfaces. All the orbital energies are defined using the vacuum as the reference state at 0 eV. (B) Schematic illustration of plasmondriven dimerization of 4-ATP chemisorbed on Ag NC surfaces under 785 nm laser excitation. (C) Time-resolved SERS spectra collected on one 4-ATP-coated SiO2@Ag NCs SNP upon exposure to 785 nm laser illumination under ambient conditions. The spectral acquisition time was 2 s and the laser power was 0.45 mW. (D) Snapshot SERS spectra collected at reaction times of 0, 4, 20, and 60 s. The highlighted 1440 cm-1, 1390 cm-1, and 1140 cm-1 bands are spectral signatures of DMAB. (E) Temporal evolution of θDMAB obtained from the time-resolved SERS results shown in panel C on one SiO2@Ag NCs SNP (left panel) and θDMAB trajectories collected on 10 different SiO2@Ag NCs SNPs under identical experimental conditions (right panel). The least-squares curve fitting results are shown as solid curve in the left panel. Plots of (F) k, and (G) θt=∞ versus I1078 cm-1 collected on individual SiO2@Ag NCs SNPs at excitation powers of 0.21, 0.32, 0.45, 0.56, and 0.90 mW. The time-resolved SERS measurements were performed on 10 different SiO2@Ag NCs SNPs at each excitation power, and the error bars represent the standard deviations of least squares curve fitting. The inset of panel F shows the plots of the ensemble averaged I1078 cm-1 as a function of laser power square, Plaser2.

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 34

We quantified the temporal evolution of the apparent fraction of DMAB, θDMAB, based on the relative peak intensities of the 1440 cm-1 Raman mode (N=N stretching mode of DMAB) with respect to the C-S stretching modes of the molecules around 1078 cm-1 (see details in Supporting Information). In the left panel of Figure 2E, we plotted the temporal evolution of θDMAB on one SiO2@Ag NCs SNP, which exhibited a typical first order kinetic curve. The apparent first order rate constant, k, and the maximum reaction yield at infinitely long reaction time, θt=∞, were obtained by performing least squares curve fitting to the θDMAB trajectory with the following first order rate equation:

 DMAB  t   (1  e k t )

(1).

We repeated the time-resolved SERS measurements on 10 different SiO2@Ag NCs SNPs oneparticle-at-a-time under exactly the same reaction conditions and all the θDMAB trajectories were shown in the right panel of Figure 2E. Although both k and θt=∞ varied from particle to particle due to the intrinsic structural heterogeneity and the variation of average field enhancements among the SNPs, all the θDMAB trajectories could be well-fitted with the first order rate law. The photocatalytic coupling of 4-ATP is mechanistically complex, involving the plasmonic excitation of the SNPs, photo-activation of interfacial oxygen, and multistep chemical transformations cascaded by multiple intermediates along the reaction pathways. However, the kinetics of DMAB formation obeyed a surprisingly simple first order rate law, strongly indicating that the overall reaction kinetics was determined by one kinetically sluggish rate-limiting step associated with the formation of a transient intermediate species that rapidly transformed into other short-lived intermediates or the final product, DMAB. The first order kinetics observed here could be most reasonably interpreted as a consequence of molecular transformations driven by photo-activated oxygen at its steady state concentrations, which could be described by two key rate-determining steps shown below. ad

Step 1:

O2

 hot e 

e





k1 k 1



O2

ACS Paragon Plus Environment

12

Page 13 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Step 2:



k

fast

2 O 2  4 ATP  TI *   1/2 DMAB

The photo-activation of interfacial oxygen involved the injection of energetic hot electrons from Ag to surface-adsorbed O2, denoted as O2ad, following the plasmonic excitation. k1 is the rate constant for the photo-activation, while k-1 is the rate constant defining the kinetics of the reverse reaction, the thermal relaxation of the photo-activated oxygen back to its ground state. In the context of the proposed steady state kinetics, both the photo-activation and thermal deactivation of interfacial oxygen (Step 1) were considered much faster than the rate-limiting step (Step 2), rapidly reaching an equilibrium that kept the photo-activated oxygen at its steady state concentration. The photo-activated oxygen subsequently reacted with the surface-adsorbed 4-ATP to initiate the stepwise oxidative dehydrogenation and azo-coupling reactions to eventually form DMAB. While the transient intermediates were unidentifiable by time-resolved SERS possibly due to their short life times, the rates of DMAB formation were essentially determined by the rate-limiting step with a rate constant of k2, which was most likely associated with the formation of a partially dehydrogenated transient intermediate, denoted as TI*. All the other faster steps along the reaction pathways, however, could be ignored in this simplified two-step mechanism, based on which we were able to derive the rate law describing the formation kinetics of DMAB. The formation rates of the photo-activated oxygen and DMAB could be described by Equations (2) and (3), respectively, where all the concentration terms referred to the interfacial abundance of the species adsorbed on photocatalyst surfaces. 

d [O2 ] ad    k1[O2 ]  k1[O2 ]  k2 [O2 ][4  ATP ] dt

(2).

d [ DMAB ]   k 2 [O2 ][4  ATP ] dt

(3). 

When the photo-activated oxygen reached its steady state concentration, [O2 ]ss ,

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 34



d [O2 ]ss   ad  k1[O2 ]  k1[O2 ]ss  k2 [O2 ]ss [4  ATP ]  0 dt

(4).

Rearrangement of the terms in Equation (4) led to the following expression ad



[O2 ]ss 

k1[O2 ] k 1  k 2 [ 4  ATP ]

(5).

Considering that the values of both k1 and k-1 were far greater than that of k2, Equation (5) could be further simplified as 

[O2 ]ss 

k1 ad [O2 ] k 1

(6).

The rate of DMAB formation in the presence of steady state O2- could be expressed as d [ DMAB ]   k 2 [O2 ]ss [4  ATP ] dt

(7).

Therefore, the apparent first-order rate constant, k, obtained from the experimentally resolved 

kinetic data was related to both k2 and [O2 ]ss as described by the following equation: 

k  k2 [O2 ]ss

(8).

Combining Equations (6) and (8) resulted in the following expression for k, k  k2

k1 ad [O2 ] k1

(9).

Equation (9) related the experimentally measured k to k1, k-1, k2, as well as the interfacial abundance ad

of surface-adsorbed 3O2, [O2 ] , providing a central guiding principle for us to rationally maneuver the rate of DMAB formation by varying the reaction conditions. While k-1 was intrinsically tied to the nature of molecular oxygen and Ag surfaces, k1 could be experimentally modulated by either adjusting the excitation laser power or varying the field enhancements of the plasmonic photocatalysts. The rate of DMAB could also be modulated by varying the partial pressure of O2 in the reaction atmosphere, ad

which controlled [O2 ] on the photocatalyst surfaces. In contrast, k2 was unrelated to the photo-

ACS Paragon Plus Environment

14

Page 15 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

activation of interfacial oxygen but varied with the molecular structures of the thiophenol derivatives as well as the photothermal annealing of the surface adsorbates on the plasmonic photocatalysts. Here ad

we demonstrated that k1, k2, and [O2 ] could all be systematically tuned to fine-maneuver the reaction kinetics by deliberately tailoring several key experimental parameters, through which compelling experimental evidences could be obtained to further verify the proposed steady state kinetic model. We first demonstrated that the rate of DMAB formation could be maneuvered by modulating k1 through variation of the excitation power while keeping all the other experimental parameters unchanged. We measured the kinetics of DMAB formation on one SiO2@Ag NCs SNP each time at 5 different excitation powers of 0.21, 0.32, 0.45, 0.56, and 0.90 mW. At each excitation power, the single-particle SERS-based kinetic measurements were performed on 10 individual SNPs. On a given SNP photocatalyst with field enhancement of |E|/|E0|, the amplitude of the enhanced plasmonic fields, |E|, was directly proportional to the electric field intensity of the incident laser, |E0|. As shown in the inset of Figure 2F, the SERS intensity of 4-ATP at 1078 cm-1 prior to the photoreactions, I1078 cm-1, was proportional to the square of the excitation power, Plaser2 (Plaser  |E0|2), further verifying the |E|4dependence of SERS signals when the Raman enhancements were dominated by the electromagnetic mechanism.4,50-52 Increasing the excitation power led to faster generation of hot electrons and higher magnitudes of local fields, both of which were highly desired for boosting the photo-activation of interfacial oxygen, resulting in higher k1 values without changing k-1. As a consequence, the equilibrium between photo-activation and thermal deactivation of interfacial oxygen was shifted toward the photo-activated state at higher excitation powers, increasing the steady state concentration of the photo-activated oxygen and thus the rate of DMAB formation (Figure 2F). The single-particle SERS results further enabled us to quantitatively correlate the reaction rates with the local fields on the surfaces of the photocatalysts. While both k and I1078

cm-1

varied form

particle to particle even at the same excitation power due to the structural heterogeneity among the

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

SiO2@Ag NCs SNPs, a general trend that k increased with I1078

cm-1

Page 16 of 34

could be clearly observed (Figure

2F). At excitation power below 0.56 mW, k was found to be proportional to I1078 cm-1. Because I1078 cm-1 scaled with the fourth power of average field intensity, |E|4, k was also proportional to |E|4. The |E|4dependence of k we observed suggested that the hot electron transfer from Ag to the 2π* orbitals of surface-adsorbed 3O2 was most likely mediated by Landau damping rather than chemical interface damping. While the rate of direct hot electron transfer during chemical interface damping is proportional to the transition dipole moment, which scales with |E|2, Landau damping involves two key steps, the excitations of hot electrons and holes in the metal and the subsequent transfer of hot electrons from the metal to the molecular adsorbates. The rates of both steps are proportional to |E|2, giving rise to the overall |E|4-dependence. As previously calculated by DFT, the oscillator strengths of plasmon excitations were much larger than that those of the direct charge transfer transitions between Ag and surface-adsorbed oxygen.48 Both our experimental observations and previous DFT calculations coherently indicated that the hot electron transfer through Landau damping was the major charge transfer channel for the photo-activation of oxygen on the surfaces of the Ag photocatalysts under near infrared excitations. Although its quantum efficiency of hot electron transfer is typically much lower than that of chemical interface damping,6,13,31 Landau damping provides more room for us to boost the photocatalytic reactions by optimizing the near-field plasmonic properties of the photocatalysts because the rates of photocatalytic reactions mediated by Landau damping are more sensitively dependent on the local fields (|E|4 for Landau damping vs. |E|2 for chemical interface damping). Whether such |E|4- vs. |E|2-dependence of k can be used as a universal criterion to distinguish Landau damping from chemical interface damping still needs to be further tested in other plasmon-driven photocatalytic reaction systems. In this work, we excited the plasmon resonance at 785 nm to ensure that the interfacial oxygen was photo-activated upon the transfer of a plasmonic hot electron from Ag to the 2π* orbitals of surface-adsorbed 3O2. Increasing the photon energy of the excitation laser allows the hot electrons to be excited to higher energies with respect to the Ag Fermi level with drastically ACS Paragon Plus Environment

16

Page 17 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

modified energy distribution profiles, making it possible to switch the dominant hot electron decay pathways from Landau damping to chemical interface damping and even to inject the hot electrons into other unoccupied orbitals of the molecular adsorbates. Detailed investigations of the effects of the excitation wavelength on the reaction kinetics and mechanisms are currently underway. As shown in Figure 2F, k exhibited a linear dependence on I1078

cm-1

in the low excitation regime

below 0.56 mW but started to show a superlinear power-law dependence on I1078

cm-1

as the excitation

powder was further increased to 0.90 mW. We hypothesized that at relatively low excitation powers, the photothermal effect was insignificant. However, above a certain threshold excitation power, the photothermal heating of the surface-adsorbed thiophenol derivatives became an additional dominant factor that further speeded up the reactions. Another noteworthy observation was that θt=∞ also increased with I1078 cm-1, in other words, θt=∞ increased with |E| on the photocatalysts. At relatively high excitation power above 0.56 mW, the reactions proceeded to θt=∞ very close to 1 on most SiO2@Ag NCs SNPs while at lower excitation powers, the majority of SiO2@Ag NCs SNPs exhibited θt=∞ significantly lower than 1 (Figure 2G). At first glance, the observed relationship between θt=∞ and |E| could be most reasonably interpreted as the characteristics of a reversible coupling/decoupling reaction. After the photocatalytic coupling reaction proceeded to a certain point, an equilibrium between the coupling and decoupling was established, reaching a non-unity θt=∞. In this case, the k obtained from the kinetic results represented the sum of the forward and reverse rate constants, while θt=∞ was determined by the ratio between forward and reverse reaction rates. Increasing the excitation power would accelerate the coupling reaction without changing the decoupling reaction rates, thereby shifting the equilibrium toward the DMAB and resulting in increased θt=∞. However, after the 4-ATP molecules were photocatalytically converted into DMAB on SiO2@Ag NCs SNPs, no reverse reaction was observed even without laser illumination. As shown in Figure S7 in Supporting Information, θt=∞ reached ~ 0.95 after illuminating a 4-ATP-coated SNPs with the 785 nm laser at 0.56 mM for ~ 50 s and no apparent decrease of θt=∞ was observed after decreasing the laser power to 0.32 mM. These ACS Paragon Plus Environment

17

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 34

observations strongly indicated that the photocatalytic coupling of 4-ATP could be considered as an irreversible reaction under the current conditions. The reverse reaction, decoupling of DMAB, was so slow that it became unobservable within the experimentally measured time scales. We hypothesized that the partial conversion of 4-ATP to DMAB at low excitation powers was due to the presence of a sub-population of surface-adsorbed 4-ATP that remained unreactive. This unreactive sub-population, however, may be thermally activated by photothermal annealing following the plamsonic excitations, which well-explained why θt=∞ increased with both the optical field, |E|, of the photocatalysts and the excitation power. The effects of the photothermal annealing will be discussed in greater detail later in this paper. As shown by Equation (9), k could also be modulated by varying the abundance of surface-adsorbed oxygen, [O2ad], without changing k1, k-1, and k2, which could be experimentally realized by adjusting the partial pressure of O2, pO2, in the reaction atmosphere. As illustrated in Figure 3A, only when the molecular O2 was physisorbed on the Ag surfaces could the plasmonic hot electrons be transferred from Ag to O2 to generate photo-activated oxygen. [O2-]ss is directly proportional to [O2ad], which was controlled by pO2. To control pO2 in the reaction atmosphere, we co-flowed O2 and N2 into the reaction chamber with controlled relative flow rates while keeping the total pressure at 1 atm, using a flow cell with an inlet and an outlet similar to the ones we previously used for single-molecule fluorescence measurements53 except that the flow cell was assembled on a silicon substrate instead of a polyethylene glycol-functionalized glass coverslip. As shown in Figures 3B and 3C, k increased with pO2, further verifying the reaction mechanism we proposed. Although kinetically very slow, the photocatalytic oxidative coupling of 4-ATP still took place with a small but non-zero k even in pure N2 atmosphere (pO2 = 0), mostly likely driven by plasmonic hot holes. As illustrated in Figure 2A, the highest occupied molecular orbital (HOMO) of 4-ATP chemisorbed on Ag was calculated to be -5.5 eV vs. vacuum, 1.2 eV below the Ag Fermi level.48,49 In principle, it is also possible to use an electron in the HOMO of chemisorbed 4-ATP to fill a hot hole in Ag following plasmonic excitation at 785 nm. ACS Paragon Plus Environment

18

Page 19 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

However, the plasmonic hot hole-driven oxidative coupling of 4-ATP was kinetically much slower than its hot electron-driven counterpart under the aerobatic conditions because of the low abundance of hot holes below -5.5 eV vs. vacuum. At 785 nm, we essentially excited the plasmonic intraband transitions, which preferentially favored the generation of energetic hot electrons exploitable for photocatalytic reactions, while efficient generation of energetic hot holes requires excitation of the interband transitions of Ag at wavelengths shorter than 330 nm.15 As shown in Figures 3B and 3D, θt=∞ also increased with pO2, suggesting that a O2-rich atmosphere was favorable to the thermal activation of surface-adsorbed molecular 4-ATP. We hypothesized that a strong synergy existed between the photo-activation of surface-adsorbed oxygen and the photothermal activation of chemisorbed 4-ATP, though the detailed mechanisms still remained open to further scrutiny.

Figure 3. Effects of interfacial oxygen abundance on photocatalytic coupling of 4-ATP. (A) Schematic illustration of the coexistence of three oxygen species: free O2 in the atmosphere, O2 adsorbed on Ag surfaces (O2ad), and photo-activated O2 (O2-). (B) θDMAB as a function of reaction time, t, under laser illumination (785 nm, 0.45 mW) at various pO2. The spectral acquisition time for each time-resolved SERS spectra was 2 s. The θDMAB trajectories show the averaged results collected on 10 different SiO2@Ag NCs SNPs under each experimental condition and the error bars represent the standard deviations. The results of least-squares fitting are shown as solid curves in panel B. (C) k and (D) θt=∞ at various pO2. The error bars show the standard deviations of least squares curve fitting. ACS Paragon Plus Environment

19

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 34

While both k1 and [O2ad] were the key factors determining [O2-]ss, k2 was intimately related to the molecular structures and interfacial behaviors of the thiophenol derivatives chemisorbed on Ag surfaces. In the context of our proposed kinetic model, k2 was determined by the activation energy barrier of the rate limiting step, which was associated with the dissociation of the N-H bonds in the amine group of 4-ATP. Therefore, when substituting the amine group of 4-ATP with a dimetheylamine group (4-DMATP) or a acetylamine group (4-AATP), k2 changed significantly because it was tied to the nature and strength of the chemical bonds on the N atom. The SERS and normal Raman spectra of 4-ATP, 4-DMATP and 4-AATP were shown in Figure S8 and the assignments of the major SERS peaks were listed in Tables S1, S3, and S4 in Supporting Information. As shown in Figure 4A, all the three thiophenol derivatives, when chemisorbed on the SiO2@Ag NCs SNPs, underwent plasmondriven oxidative coupling processes to form DMAB upon exposure to 785 nm laser at an excitation power of 0.90 mW. Analogous to that of 4-ATP, the photocatalytic coupling reactions of both 4DMATP and 4-AATP also obeyed the first-order rate law (Figures S9 and S10 in Supporting Information). The first-order rate constants of both the 4-DMATP and 4-AATP coupling reactions were observed to be proportional to initial SERS intensity of the C-S stretching modes (1082 cm-1 for 4-DMATP and 1075 cm-1 for 4-AATP) in low excitation power regimes and switched to a superlinear power-law dependence at higher excitation powers (upper panels in Figures 4B and 4C). Similar to that of 4-ATP coupling, the θt=∞ for the 4-DMATP and 4-AATP coupling reactions also increased with initial SERS intensity of the C-S stretching modes, approaching 100 % apparent reaction yields at sufficiently high excitation powers (lower panels in Figures 4B and 4C). These results strongly indicated that the proposed mechanism for the photocatalytic oxidative coupling of 4-ATP, which involved the steady state photo-activated oxygen, could be used as a generic kinetic model to interpret the kinetics of 4-DMATP and 4-AATP coupling reactions as well. However, both k and θt=∞ varied

ACS Paragon Plus Environment

20

Page 21 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

significantly under the same experimental conditions (the same excitation power and pO2) when the molecular structures of the thiophenol derivatives changed. As shown in Figures 4D and 4E, both k and θt=∞ followed the same trend of 4-ATP > 4-DMATP > 4-AATP.

Figure 4. Molecular structural effects on photocatalytic coupling of thiophenol derivatives. (A) Representative SERS spectra of 4-ATP, 4-DMATP, and 4-AATP (i) before and (ii) after the oxidative coupling reaction under laser illumination (785 nm, 0.90 mW) for 120 s. The spectra acquisition time was 1 s for 4-ATP, and 2 s for 4-DMATP and 4-AATP, respectively. (B) Plots of k (upper panel) and θt=∞ (lower panel) versus the initial SERS intensities at 1082 cm-1, I1082 cm-1, collected on individual 4DMATP-coated SiO2@Ag NCs SNPs at excitation powers of 0.32, 0.56, 0.90, 1.40, and 2.60 mW. (C) Plots of k (upper panel) and θt=∞ (lower panel) versus the initial SERS intensities at 1075 cm-1, I1075 cm-1, collected on individual 4-AATP-coated SiO2@Ag NCs SNPs at excitation powers of 0.56, 0.90, 2.00, 2.60, and 5.00 mW. The time-resolved SERS measurements were performed on 10 different SiO2@Ag NCs SNPs at each excitation power, and the error bars represented the standard deviations of least squares curve fitting. Comparison of (D) k and (E) θt=∞ of 4-ATP, 4-DMATP, and 4-AATP at excitation powers of 0.56 mW and 0.90 mW. The error bars represented the standard deviations of 10 SiO2@Ag NCs SNPs under each condition.

The different k values among the three thiophenol derivatives essentially stemmed from the difference in k2, which was related to the polarities and dissociation energies of the chemical bonds. k2 ACS Paragon Plus Environment

21

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 34

for 4-ATP was significantly higher than that for 4-DMATP under the same experimental conditions because the N-H bond in 4-ATP was remarkably more polar and thus easier to dissociate than the N-C bond in 4-DMATP when reacting with the photo-activated oxygen. In addition, 4-ATP exhibited smaller steric hindrance than 4-DMATP for the photo-activated oxygen to attack the leaving groups, which were H atoms for 4-ATP and methyl groups for 4-DMATP, respectively. Among the three thiophenol derivatives, 4-AATP exhibited the smallest k and lowest θt=∞ because the N-C bond in 4AATP was partially double-bonded in nature as a consequence of electron delocalization in the acetylamine group, thereby creating a markedly higher energy barrier for the bond dissociation. k2 was found to be related to not only the intrinsic molecular structures but the photothermal heating of the surface-adsorbed thiophenol derivatives as well. Following the optical excitation of plasmon resonances, the energetic hot electrons may also undergo thermal dissipation to heat up the metal lattices via electron-phonon coupling.10 At relatively low excitation power, the photothermal heating of the interfacial molecular adsorbates had insignificant impact to the overall reaction kinetics. Therefore, k was observed to be proportional to |E|4 of the photocatalysts for all 4-ATP (Figure 2F), 4-DMATP (Figure 4B), and 4-AATP (Figure 4C). Above a certain threshold excitation power, however, a superlinear power-law dependence of k on the |E|4 of the photocatalysts was observed due to increased local reaction temperatures and thermally induced restructuring of the molecular adsorbates, both of which led to decrease of reaction energy barriers, and thus increased k2. In Figure S11 in Supporting Information, we compared the SERS spectra of 4-ATP, 4-DMATP, and 4-AATP before and after laser illumination for 60 seconds in the presence of pure N2. While almost no DMAB formation was observed within 60 seconds in such a O2-free environment, significant increase in the SERS peak intensities was clearly observed on all the three thiophenol derivatives, which could be interpreted as a consequence of photothermally induced strengthening of the interactions between the adsorbate molecules and Ag NC surfaces. We also observed that the SERS intensities of 4-ATP increased by about 100% upon thermal annealing at 90 ⁰C for 1 hour prior to SERS measurements (Figure S12 in ACS Paragon Plus Environment

22

Page 23 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Supporting Information). The additional SERS enhancements upon thermal annealing was also previously observed by Aggarwal and coworkers on the benzenethiol adsorbed on nanostructured Ag films, which was interpreted as a consequence of thermally induced transitions between various adsorption states of benzenethiol.54 To further verify the photothermal effects on the reaction rate, we measured the kinetics of the photocatalytic coupling of 4-ATP with time intervals of 60 seconds inserted between successive SERS spectral acquisitions instead of continuous illumination (Figure S13 in Supporting Information). Adding the time intervals allowed the photothermally generated heat to dissipate to the environment before the next cycle of laser excitation started such that the local heating at the molecule-nanoparticle interfaces could be minimized. Interestingly, switching from continuous to discontinuous illuminations did not result in much difference in k and I1078 cm-1 at excitation power of 0.45 mW, suggesting that the photothermal effect was insignificant at this excitation power, in agreement with the results shown in Figure 2F. However, obvious decrease in both k and I1078 cm-1 was observed when the time intervals were included during the kinetic measurements at the excitation power of 0.90 mW, which was in a power regime where k exhibited a superlinear power-law dependence on |E|4 due to significant photothermal effects. The plasmonic photothermal annealing of the surface-adsorbed thiophenol derivatives also greatly influenced θt=∞. In the low excitation power regime, only a fraction of 4-ATP molecules, most likely the ones experiencing the strongest fields in the interstitial plasmonic hot spots, could be photothermally activated and subsequently transformed into DMAB, while those in the plasmonic “colder” region remained unreactive due to insufficient photothermal heating. Takeyasu and co-authors recently discovered that only when the excitation laser power was above a threshold value could the photocatalytic coupling of 4-ATP occur,55 though the underlying mechanisms were not further discussed. Although the hot spots provided the locations for the strongest SERS enhancements, the

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 34

unreactive 4-ATP in the plasmonic “colder” regions also contributed to the overall SERS signals. As the excitation power increased, increasing fractions of the molecules started to be located in regions with local fields sufficiently high for the thermal activation, resulting in increased θt=∞ until θt=∞ approached 1 at sufficiently high excitation powers, nominally 0.56 mW for 4-ATP, 1.40 mW for 4DMATP, and 2.60 mW for 4-AATP, respectively, under the current experimental conditions.

Figure 5. Thermal effects on photocatalytic coupling of thiophenol derivatives. Plots of k (top panels) and θt=∞ (bottom panels) versus the initial SERS intensities of the C-S stretching modes of (A) 4-ATP, (B) 4-DMATP, and (C) 4-AATP. The kinetic results obtained on pre-annealed samples and those without pre-annealing treatments were compared under various excitation laser intensities. Under each condition, the time-resolved SERS measurements were performed on 10 different SiO2@Ag NCs SNPs and the error bars represented the standard deviations of least squares curve fitting.

The thermal effects on the kinetics and yields of the photocatalytic oxidative coupling of 4-ATP, 4DMATP, and 4-AATP were further verified by pre-annealing the samples at 90 ⁰C for 1 h to optimize the structures and thus thermally activate the surface-adsorbed thiophenol derivatives prior to the timeresolved SERS measurements. In Figure 5, we showed the correlations of k and θt=∞ with the initial SERS intensities of the C-S stretching modes of the reactants for both pre-annealed and unannealed ACS Paragon Plus Environment

24

Page 25 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

samples. The pre-annealing of the samples gave rise to further enhanced SERS signals while the positions and the lineshapes of all the characteristic SERS peaks were well-preserved, verifying that the thermal treatments of the samples neither induced any thermal coupling reactions nor caused any sample damage. The pre-annealing treatments resulted in significant increase of both k and θt=∞ for all the three thiophenol derivatives we investigated in comparison to those of the unannealed samples, and this generic trend held true when varying the excitation power. The results shown in Figure 5 provided important insights into the underlying thermal effects on the kinetics and yields of the photocatalytic oxidative coupling of thiophenol derivatives, which well-interpreted the superlinear power-law dependence of k and θt=∞ on the local fields of the photocatalysts observed at high excitation powers.

CONCLUSIONS We have resolved the complex kinetics of photocatalytic oxidative coupling of thiophenol derivatives chemisorbed on the surfaces of optically excited plasmonic photocatalysts using timeresolved SERS as a plasmon-enhanced spectroscopic tool. This work provides quantitative mechanistic insights on a series of fundamentally important issues that have long been poorly understood. First, the rate limiting steps have been identified to be associated with the thermal reactions between photoactivated oxygen and surface-adsorbed thiophenol derivatives rather than the plasmon-driven photoactivation of interfacial oxygen. Second, photo-activation of interfacial oxygen occurs upon transfer of a hot electron from Ag to the physisorbed oxygen through Landau damping. When the photoactivation of oxygen is much faster than the rate limiting step, the rates of the coupling reactions are dynamically maneuvered by the photo-activated oxygen at its steady state concentrations, giving rise to an apparent first-order overall reaction kinetics. The steady state concentration of photo-activated oxygen can be experimentally tuned by selectively varying the excitation power, the local optical fields, and the abundance of interfacial oxygen. Third, time-resolved SERS measurements allow us to ACS Paragon Plus Environment

25

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 34

quantitatively correlate the reaction rates with the local field intensities on the surfaces of the photocatalysts. The rate constants switch from a linear dependence to a superlinear power-law dependence with respect to the fourth power of the local fields as the excitation power progressively increases essentially due to the plasmonic photothermal effects. Fourth, the rate constants of the rate limiting steps are related to the molecular structures and thermal restructuring of the thiophenol derivatives chemisorbed on the nanoparticle surfaces. The plamsonic photothermal heating of the surface-adsorbates may not only modify the molecular adsorption states but also thermally activate the thiophenol derivatives for the subsequent oxidative coupling reactions, significantly enhancing both the rates and the yields of the reactions. The knowledge gained from this work provides key design principles guiding the rational optimization of plasmon-driven photocatalytic processes. The photo-activated oxygen has been found to be a crucial transient species broadly involved in a variety of photocatalytic oxidation reactions.11,14,30-34 Therefore, the insights gained from the time-resolved SERS measurements may serve as a generic central knowledge framework that guides us to rationally maneuver the kinetics of a wide range of oxidation reactions driven by energetic hot electrons. Analogous to oxygen, interfacial molecular hydrogen can also be photo-activated when interacting with optically excited plasmonic nanostructures.23,24 We believe that the plasmonically activated hydrogen species may be harnessed to drive photocatalytic reductive chemical transformations as well. The kinetic model involving the steady state photo-activated species proposed in this work provides important implications that may broadly apply to a large variety of plasmon-driven photocatalytic reactions, greatly enhancing our capabilities to selectively and efficiently drive unconventional photocatalytic chemical transformations by fine-tailoring the plasmonic characteristics of the photocatalysts and the interfacial moleculenanoparticle interactions.

ACS Paragon Plus Environment

26

Page 27 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

ASSOCIATED CONTENT Supporting Information

Additional experimental details, tables listing Raman peak assignments, and figures including SEM images, TEM images, dark-field scattering images, ζ-potential results, and time-resolved SERS results as noted in the text. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author

*Email: [email protected] (H.W.); Phone: 803-777-2203; Fax: 803-777-9521.

Notes

The authors declare no competing financial interests.

ACKNOWLEDGMENT This work was supported by a seed grant provided by the University of South Carolina (USC) Office of Vice President for Research through an ASPIRE-I Track-I Award and in part by National Science Foundation through a CAREER Award (DMR-1253231) and an EPSCoR RII Track-I Award (OIA1655740). Q.Z. was partially supported by a Dissertation Fellowship from USC NanoCenter and a SPARC Graduate Research Award from the USC Office of the Vice President for Research. This work made use of the facilities at the USC Electron Microscopy Center and W.M. Keck Open Laboratory.

ACS Paragon Plus Environment

27

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 34

REFERENCES (1)

Schuller, J. A.; Barnard, E. S.; Cai, W. S.; Jun, Y. C.; White, J. S.; Brongersma, M. L.

Plasmonics for Extreme Light Concentration and Manipulation. Nat. Mater. 2010, 9, 193-204. (2)

Kauranen, M.; Zayats, A. V. Nonlinear Plasmonics. Nat. Photonics 2012, 6, 737-748.

(3)

Ciraci, C.; Hill, R. T.; Mock, J. J.; Urzhumov, Y.; Fernandez-Dominguez, A. I.; Maier,

S. A.; Pendry, J. B.; Chilkoti, A.; Smith, D. R. Probing the Ultimate Limits of Plasmonic Enhancement. Science 2012, 337, 1072-1074. (4)

Fang, Y.; Seong, N. H.; Dlott, D. D. Measurement of the Distribution of Site

Enhancements in Surface-Enhanced Raman Scattering. Science 2008, 321, 388-392. (5)

Knight, M. W.; Sobhani, H.; Nordlander, P.; Halas, N. J. Photodetection with Active

Optical Antennas. Science 2011, 332, 702-704. (6)

Wu, K.; Chen, J.; McBride, J. R.; Lian, T. Efficient Hot-Electron Transfer by a

Plasmon-Induced Interfacial Charge-Transfer Transition. Science 2015, 349, 632-635. (7)

Clavero, C. Plasmon-Induced Hot-Electron Generation at Nanoparticle/Metal-Oxide

Interfaces for Photovoltaic and Photocatalytic Devices. Nat. Photonics 2014, 8, 95-103. (8)

Manjavacas, A.; Liu, J. G.; Kulkarni, V.; Nordlander, P. Plasmon-Induced Hot Carriers

in Metallic Nanoparticles. ACS Nano 2014, 8, 7630-7638. (9)

Govorov, A. O.; Zhang, H.; Gun'ko, Y. K. Theory of Photoinjection of Hot Plasmonic

Carriers from Metal Nanostructures into Semiconductors and Surface Molecules. J. Phys. Chem. C 2013, 117, 16616-16631.

(10)

Brongersma, M. L.; Halas, N. J.; Nordlander, P. Plasmon-Induced Hot Carrier Science

and Technology. Nat. Nanotechnol. 2015, 10, 25-34. (11)

Christopher, P.; Xin, H. L.; Marimuthu, A.; Linic, S. Singular Characteristics and

Unique Chemical Bond Activation Mechanisms of Photocatalytic Reactions on Plasmonic Nanostructures. Nat. Mater. 2012, 11, 1044-1050. ACS Paragon Plus Environment

28

Page 29 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(12)

Mubeen, S.; Lee, J.; Singh, N.; Kramer, S.; Stucky, G. D.; Moskovits, M. An

Autonomous Photosynthetic Device in Which All Charge Carriers Derive from Surface Plasmons. Nat. Nanotechnol. 2013, 8, 247-251. (13)

Linic, S.; Aslam, U.; Boerigter, C.; Morabito, M. Photochemical Transformations on

Plasmonic Metal Nanoparticles. Nat. Mater. 2015, 14, 567-576. (14)

Linic, S.; Christopher, P.; Ingram, D. B. Plasmonic Metal Nanostructures for Efficient

Conversion of Solar to Chemical Energy. Nat. Mater. 2011, 10, 911-921. (15)

Govorov, A. O.; Zhang, H.; Demir, H. V.; Gun'ko, Y. K. Photogeneration of Hot

Plasmonic Electrons with Metal Nanocrystals: Quantum Description and Potential Applications. Nano Today 2014, 9, 85-101. (16)

Brus, L. Noble Metal Nanocrystals: Plasmon Electron Transfer Photochemistry and

Single-Molecule Raman Spectroscopy. Acc. Chem. Res. 2008, 41, 1742-1749. (17)

Jin, R. C.; Cao, Y. W.; Mirkin, C. A.; Kelly, K. L.; Schatz, G. C.; Zheng, J. G.

Photoinduced Conversion of Silver Nanospheres to Nanoprisms. Science 2001, 294, 1901-1903. (18)

Jin, R. C.; Cao, Y. C.; Hao, E. C.; Metraux, G. S.; Schatz, G. C.; Mirkin, C. A.

Controlling Anisotropic Nanoparticle Growth through Plasmon Excitation. Nature 2003, 425, 487-490. (19)

Marimuthu, A.; Zhang, J. W.; Linic, S. Tuning Selectivity in Propylene Epoxidation by

Plasmon Mediated Photo-Switching of Cu Oxidation State. Science 2013, 339, 1590-1593. (20)

Zhai, Y. M.; DuChene, J. S.; Wang, Y. C.; Qiu, J. J.; Johnston-Peck, A. C.; You, B.;

Guo, W. X.; DiCiaccio, B.; Qian, K.; Zhao, E. W., et al. Polyvinylpyrrolidone-Induced Anisotropic Growth of Gold Nanoprisms in Plasmon-Driven Synthesis. Nat. Mater. 2016, 15, 889-895. (21)

Deeb, C.; Ecoffet, C.; Bachelot, R.; Plain, J.; Bouhelier, A.; Soppera, O. Plasmon-Based

Free-Radical Photopolymerization: Effect of Diffusion on Nanolithography Processes. J. Am. Chem. Soc. 2011, 133, 10535-10542.

ACS Paragon Plus Environment

29

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(22)

Page 30 of 34

Ueno, K.; Juodkazis, S.; Shibuya, T.; Yokota, Y.; Mizeikis, V.; Sasaki, K.; Misawa, H.

Nanoparticle Plasmon-Assisted Two-Photon Polymerization Induced by Incoherent Excitation Source. J. Am. Chem. Soc. 2008, 130, 6928-6929. (23)

Mukherjee, S.; Libisch, F.; Large, N.; Neumann, O.; Brown, L. V.; Cheng, J.; Lassiter,

J. B.; Carter, E. A.; Nordlander, P.; Halas, N. J. Hot Electrons Do the Impossible: Plasmon-Induced Dissociation of H2 on Au. Nano Lett. 2013, 13, 240-247. (24)

Mukherjee, S.; Zhou, L. A.; Goodman, A. M.; Large, N.; Ayala-Orozco, C.; Zhang, Y.;

Nordlander, P.; Halas, N. J. Hot-Electron-Induced Dissociation of H2 on Gold Nanoparticles Supported on SiO2. J. Am. Chem. Soc. 2014, 136, 64-67. (25)

Lee, J.; Mubeen, S.; Ji, X. L.; Stucky, G. D.; Moskovits, M. Plasmonic Photoanodes for

Solar Water Splitting with Visible Light. Nano Lett. 2012, 12, 5014-5019. (26)

Xie, W.; Schlucker, S. Hot Electron-Induced Reduction of Small Molecules on

Photorecycling Metal Surfaces. Nat. Commun. 2015, 6, 7570. (27)

Brandt, N. C.; Keller, E. L.; Frontiera, R. R. Ultrafast Surface-Enhanced Raman

Probing of the Role of Hot Electrons in Plasmon-Driven Chemistry. J. Phys. Chem. Lett. 2016, 7, 3179-3185. (28)

Sun, M. T.; Xu, H. X. A Novel Application of Plasmonics: Plasmon-Driven Surface-

Catalyzed Reactions. Small 2012, 8, 2777-2786. (29)

van Schrojenstein Lantman, E. M.; Deckert-Gaudig, T.; Mank, A. J. G.; Deckert, V.;

Weckhuysen, B. M. Catalytic Processes Monitored at the Nanoscale with Tip-Enhanced Raman Spectroscopy. Nat. Nanotechnol. 2012, 7, 583-586. (30)

Christopher, P.; Xin, H. L.; Linic, S. Visible-Light-Enhanced Catalytic Oxidation

Reactions on Plasmonic Silver Nanostructures. Nat. Chem. 2011, 3, 467-472. (31)

Kale, M. J.; Avanesian, T.; Christopher, P. Direct Photocatalysis by Plasmonic

Nanostructures. ACS Catal. 2014, 4, 116-128. ACS Paragon Plus Environment

30

Page 31 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(32)

Fang, Y. R.; Li, Y. Z.; Xu, H. X.; Sun, M. T. Ascertaining p,p '-Dimercaptoazobenzene

Produced from p-Aminothiophenol by Selective Catalytic Coupling Reaction on Silver Nanoparticles. Langmuir 2010, 26, 7737-7746. (33)

Huang, Y. F.; Zhu, H. P.; Liu, G. K.; Wu, D. Y.; Ren, B.; Tian, Z. Q. When the Signal

Is Not from the Original Molecule To Be Detected: Chemical Transformation of paraAminothiophenol on Ag during the SERS Measurement. J. Am. Chem. Soc. 2010, 132, 9244-9246. (34)

Linic, S.; Christopher, P.; Xin, H. L.; Marimuthu, A. Catalytic and Photocatalytic

Transformations on Metal Nanoparticles with Targeted Geometric and Plasmonic Properties. Acc. Chem. Res. 2013, 46, 1890-1899. (35)

Wang, F.; Li, C.; Chen, H.; Jiang, R.; Sun, L.-D.; Li, Q.; Wang, J.; Yu, J. C.; Yan, C.-H.

Plasmonic Harvesting of Light Energy for Suzuki Coupling Reactions. J. Am. Chem. Soc. 2013, 135, 5588-5601. (36)

Sarina, S.; Zhu, H.; Jaatinen, E.; Xiao, Q.; Liu, H.; Jia, J.; Chen, C.; Zhao, J. Enhancing

Catalytic Performance of Palladium in Gold and Palladium Alloy Nanoparticles for Organic Synthesis Reactions through Visible Light Irradiation at Ambient Temperatures. J. Am. Chem. Soc. 2013, 135, 5793-5801. (37)

Guo, J.; Zhang, Y.; Shi, L.; Zhu, Y.; Mideksa, M. F.; Hou, K.; Zhao, W.; Wang, D.;

Zhao, M.; Zhang, X., et al. Boosting Hot Electrons in Hetero-Superstructures for Plasmon-Enhanced Catalysis. J. Am. Chem. Soc. 2017, 139, 17964-17972. (38)

Huang, Y. F.; Zhang, M.; Zhao, L. B.; Feng, J. M.; Wu, D. Y.; Ren, B.; Tian, Z. Q.

Activation of Oxygen on Gold and Silver Nanoparticles Assisted by Surface Plasmon Resonances. Angew. Chem. Int. Ed. 2014, 53, 2353-2357. (39)

Osawa, M.; Matsuda, N.; Yoshii, K.; Uchida, I. Charge-Transfer Resonance Raman

Process in Surface-Enhanced Raman-Scattering from P-Aminothiophenol Adsorbed on Silver Herzberg-Teller Contribution. J. Phys. Chem. 1994, 98, 12702-12707. ACS Paragon Plus Environment

31

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(40)

Page 32 of 34

Xu, P.; Kang, L. L.; Mack, N. H.; Schanze, K. S.; Han, X. J.; Wang, H. L. Mechanistic

Understanding of Surface Plasmon Assisted Catalysis on a Single Particle: Cyclic Redox of 4Aminothiophenol. Sci. Rep. 2013, 3, 2997. (41)

Dai, Z. G.; Xiao, X. H.; Wu, W.; Zhang, Y. P.; Liao, L.; Guo, S. S.; Ying, J. J.; Shan, C.

X.; Sun, M. T.; Jiang, C. Z. Plasmon-Driven Reaction Controlled by the Number of Graphene Layers and Localized Surface Plasmon Distribution during Optical Excitation. Light-Sci. Appl. 2015, 4, e342. (42)

Sun, M. T.; Huang, Y. Z.; Xia, L. X.; Chen, X. W.; Xu, H. X. The pH-Controlled

Plasmon-Assisted

Surface

Photocatalysis

Reaction

of

4-Aminothiophenol

to

p,p'-

Dimercaptoazobenzene on Au, Ag, and Cu Colloids. J. Phys. Chem. C 2011, 115, 9629-9636. (43)

Zheng, T. T.; Zhang, Q. F.; Feng, S.; Zhu, J. J.; Wang, Q.; Wang, H. Robust

Nonenzymatic Hybrid Nanoelectrocatalysts for Signal Amplification toward Ultrasensitive Electrochemical Cytosensing. J. Am. Chem. Soc. 2014, 136, 2288-2291. (44)

Wang, H.; Halas, N. J. Mesoscopic Au "Meatball" particles. Adv. Mater. 2008, 20, 820-

(45)

Xia, X. H.; Zeng, J.; Oetjen, L. K.; Li, Q. G.; Xia, Y. N. Quantitative Analysis of the

825.

Role Played by Poly(vinylpyrrolidone) in Seed-Mediated Growth of Ag Nanocrystals. J. Am. Chem. Soc. 2012, 134, 1793-1801. (46)

Halas, N. J.; Lal, S.; Chang, W. S.; Link, S.; Nordlander, P. Plasmons in Strongly

Coupled Metallic Nanostructures. Chem. Rev. 2011, 111, 3913-3961. (47)

Malynych, S.; Luzinov, I.; Chumanov, G. Poly(vinyl pyridine) as a Universal Surface

Modifier for Immobilization of Nanoparticles. J. Phys. Chem. B 2002, 106, 1280-1285. (48)

Zhao, L. B.; Liu, X. X.; Zhang, M.; Liu, Z. F.; Wu, D. Y.; Tian, Z. Q. Surface Plasmon

Catalytic Aerobic Oxidation of Aromatic Amines in Metal/Molecule/Metal Junctions. J. Phys. Chem. C 2016, 120, 944-955.

ACS Paragon Plus Environment

32

Page 33 of 34 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(49)

Zhao, L. B.; Zhang, M.; Huang, Y. F.; Williams, C. T.; Wu, D. Y.; Ren, B.; Tian, Z. Q.

Theoretical Study of Plasmon-Enhanced Surface Catalytic Coupling Reactions of Aromatic Amines and Nitro Compounds. J. Phys. Chem. Lett. 2014, 5, 1259-1266. (50)

Campion, A.; Kambhampati, P. Surface-enhanced Raman scattering. Chem. Soc. Rev.

1998, 27, 241-250.

(51)

Willets, K. A.; Van Duyne, R. P. Localized Surface Plasmon Resonance Spectroscopy

and Sensing. Annu. Rev. Phys. Chem. 2007, 58, 267-297. (52)

Li, J. F.; Huang, Y. F.; Ding, Y.; Yang, Z. L.; Li, S. B.; Zhou, X. S.; Fan, F. R.; Zhang,

W.; Zhou, Z. Y.; Wu, D. Y., et al. Shell-Isolated Nanoparticle-Enhanced Raman Spectroscopy. Nature 2010, 464, 392-395.

(53)

Wang, H.; Musier-Forsyth, K.; Falk, C.; Barbara, P. F. Single-Molecule Spectroscopic

Study of Dynamic Nanoscale DNA Bending Behavior of HIV-1 Nucleocapsid Protein. J. Phys. Chem. B 2013, 117, 4183-4196. (54)

Aggarwal, R. L.; Farrar, L. W.; Saikin, S. K. Increase of SERS Signal upon Heating or

Exposure to a High-Intensity Laser Field: Benzenethiol on an AgFON Substrate. J. Phys. Chem. C 2012, 116, 16656-16659.

(55)

Takeyasu, N.; Kagawa, R.; Sakata, K.; Kaneta, T. Laser Power Threshold of Chemical

Transformation on Highly Uniform Plasmonic and Catalytic Nanosurface. J. Phys. Chem. C 2016, 120, 12163-12169.

ACS Paragon Plus Environment

33

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 34

TOC Graphic:

ACS Paragon Plus Environment

34